• Non ci sono risultati.

Big Bang 6Li nucleosynthesis studied deep underground (LUNA collaboration)

N/A
N/A
Protected

Academic year: 2021

Condividi "Big Bang 6Li nucleosynthesis studied deep underground (LUNA collaboration)"

Copied!
10
0
0

Testo completo

(1)

24 July 2021

AperTO - Archivio Istituzionale Open Access dell'Università di Torino

Original Citation:

Big Bang 6Li nucleosynthesis studied deep underground (LUNA collaboration)

Published version:

DOI:10.1016/j.astropartphys.2017.01.007

Terms of use:

Open Access

(Article begins on next page)

Anyone can freely access the full text of works made available as "Open Access". Works made available under a

Creative Commons license can be used according to the terms and conditions of said license. Use of all other works

requires consent of the right holder (author or publisher) if not exempted from copyright protection by the applicable law.

Availability:

This is the author's manuscript

(2)

Contents

lists

available

at

ScienceDirect

Astroparticle

Physics

journal

homepage:

www.elsevier.com/locate/astropartphys

Big

Bang

6

Li

nucleosynthesis

studied

deep

underground

(LUNA

collaboration)

D.

Trezzi

a

,

M.

Anders

b

,

c

,

1

,

M.

Aliotta

d

,

A.

Bellini

e

,

D.

Bemmerer

b

,

A.

Boeltzig

f

,

g

,

C.

Broggini

h

,

C.G.

Bruno

d

,

A.

Caciolli

h

,

i

,

F.

Cavanna

e

,

P.

Corvisiero

e

,

H.

Costantini

e

,

2

,

T.

Davinson

d

,

R.

Depalo

h

,

i

,

Z.

Elekes

b

,

M.

Erhard

h

,

F.

Ferraro

e

,

A.

Formicola

f

,

Zs.

Fülop

j

,

G.

Gervino

k

,

A.

Guglielmetti

a

,

C.

Gustavino

l

,

,

Gy.

Gyürky

j

,

M.

Junker

f

,

A.

Lemut

e

,

3

,

M.

Marta

b

,

4

,

C.

Mazzocchi

a

,

5

,

R.

Menegazzo

h

,

V.

Mossa

m

,

F.

Pantaleo

m

,

P.

Prati

e

,

C.

Rossi

Alvarez

h

,

D.A.

Scott

d

,

E.

Somorjai

j

,

O.

Straniero

n

,

o

,

T.

Szücs

j

,

M.

Takacs

b

a Università degli Studi di Milano and INFN, Sezione di Milano, Via Celoria 16, 20133 Milano, Italy b Helmholtz-Zentrum Dresden-Rossendorf, Bautzner Landstr. 400, 01328 Dresden, Germany c Technische Universität Dresden, Mommsenstrasse 9, 01069 Dresden, Germany

d SUPA, School of Physics and Astronomy, University of Edinburgh, EH9 3JZ Edinburgh, United Kingdom e Università degli Studi di Genova and INFN, Sezione di Genova, Via Dodecaneso 33, 16146 Genova, Italy f Laboratori Nazionali del Gran Sasso (LNGS), Via Acitelli 22, 67010 Assergi (AQ), Italy

g Gran Sasso Science Institute (GSSI), Viale F. Crispi 7, 67100 L’Aquila (AQ), Italy h INFN, Sezione di Padova, Via F. Marzolo 8, 35131 Padova, Italy

i Dipartimento di Fisica e Astronomia, Università degli studi di Padova, Via F. Marzolo 8, 35131 Padova, Italy j Institute for Nuclear Research (MTA ATOMKI), PO Box 51, HU-4001 Debrecen, Hungary

k Università degli Studi di Torino and INFN, Sezione di Torino, Via P. Giuria 1, 10125 Torino, Italy l INFN, Sezione di Roma Sapienza, Piazzale A. Moro 2, 00185 Roma, Italy

m Dipartimento Interateneo, Università degli Studi di Bari and INFN, Sezione di Bari, Via E. Orabona 4, 70125 Bari, Italy n Osservatorio Astronomico di Collurania, Via M. Maggini, 64100 Teramo, Italy

o INFN, Sezione di Napoli, Via Cintia, 80126 Napoli, Italy

a

r

t

i

c

l

e

i

n

f

o

Article history:

Received 20 November 2015 Revised 2 December 2016 Accepted 22 January 2017 Available online 23 January 2017

Keywords:

Nuclear Astrophysics Big Bang Nucleosynthesis 6Li abundance Underground

a

b

s

t

r

a

c

t

Thecorrectpredictionoftheabundancesofthelightnuclides producedduringtheepochofBigBang Nucleosynthesis (BBN) is one ofthe main topics ofmodern cosmology. For manyof the nuclear re-actions thatarerelevant forthisepoch, directexperimental crosssection dataareavailable, ushering the so-called“ageof precision”.The present work addresses an exception tothis currentstatus: the 2 H(

α

,

γ

)6 Lireactionthatcontrols6 LiproductionintheBigBang.Recentcontroversialobservationsof6 Li inmetal-poorstarshaveheightenedtheinterestinunderstandingprimordial6 Liproduction.Ifconfirmed, theseobservationswouldleadtoasecondcosmologicallithiumproblem,inadditiontothewell-known 7 Liproblem. Inthepresentwork,thedirectexperimentalcrosssection dataon2 H(

α

,

γ

)6 LiintheBBN energyrangearereported.ThemeasurementhasbeenperformeddeepundergroundattheLUNA (Lab-oratoryforUndergroundNuclearAstrophysics)400kVacceleratorintheLaboratoriNazionalidelGran Sasso,Italy.ThecrosssectionhasbeendirectlymeasuredattheenergiesofinterestforBigBang Nucle-osynthesisforthefirsttime,atEcm =80,93,120,and133keV.Basedonthenew data,the 2 H(

α

,

γ

)6 Li thermonuclearreactionratehasbeenderived.Ourrateisevenlowerthanpreviouslyreported,thus in-creasingthediscrepancybetweenpredictedBigBang6 Liabundanceand theamountofprimordial6 Li inferredfromobservations.

© 2017 Elsevier B.V. All rights reserved.

Corresponding author.

E-mail address: carlo.gustavino@roma1.infn.it (C. Gustavino).

1 Present address: Sächsisches Staatministerium für Umwelt und Landwirtschaft,

Dresden, Germany.

2 Present address: CPPM, Université d’Aix-Marseille, CNRS/IN2P3, Marseille,

France.

3 Deceased.

4 Present address: GSI, Darmstadt, Germany.

5 Present address: University of Warsaw, Warsaw, Poland. http://dx.doi.org/10.1016/j.astropartphys.2017.01.007

(3)

58 D. Trezzi et al. / Astroparticle Physics 89 (2017) 57–65

1.

Introduction

Big

Bang

Nucleosynthesis

(BBN)

may

be

used

to

probe

cosmo-logical

models

and

parameters.

To

this

end,

abundance

predictions

from

BBN

have

to

be

compared

with

the

abundances

inferred

by

astronomers

observing

the

emission

and

absorption

lines

in

spe-cific

astrophysical

environments.

The

pillars

of

BBN

calculations

are:

Standard

Cosmological

Big

Bang

model,

the

Standard

Model

of

Particle

Physics,

and

the

nuclear

cross

sections

of

the

processes

involved

in

the

BBN

reaction

network.

The

agreement

between

cal-culations

and

observations

for

primordial

deuterium

and

4

He

(see

for

example

[1]

)

places

cosmology,

nuclear

and

particle

physics

in

a

uniquely

consistent

framework.

However,

the

situation

is

not

as

favorable

for

7

Li

and

6

Li

[2]

.

For

7

Li,

the

predicted

abundance

is

about

a

factor

of

3

higher

than

the

observed

one.

As

of

today

a

clear

solution

to

this

puzzle,

“the

lithium

problem

”,

has

not

been

found.

Even

more

complex

is

the

case

of

6

Li,

where

the

predicted

abundance

is

up

to

three

orders

of

magnitude

lower

than

that

in-ferred

from

direct

observations

(the

so

called

“second

lithium

prob-lem

”)

[3,4]

.

The

aim

of

the

present

work

is

to

put

the

nuclear

physics

of

6

Li

production

in

standard

Big

Bang

scenarios

on

solid

experimental

ground.

The

nuclear

cross

section

of

the

leading

process

in

the

6

Li

primordial

production,

i.e.

the

2

H(

α

,

γ

)

6

Li

fusion

reaction,

has

thus

been

directly

measured

in

the

BBN

energy

range

at

LUNA

(Labo-ratory

for

Underground

Nuclear

Astrophysics).

The

measurement

has

been

carried

out

with

the

world’s

only

underground

acceler-ator

for

nuclear

astrophysics,

situated

at

the

Laboratori

Nazionali

del

Gran

Sasso

(LNGS),

Italy.

The

data

at

center-of-mass

energy

E

=

133

and

93

keV

have

been

previously

published

in

abbreviated

form

[5]

.

Here,

two

new

data

points

at

E

=

120

and

80

keV

are

presented.

The

present

work

is

organized

as

follows:

In

Sections

2

and

3

a

description

of

astronomical

observations

and

nuclear

pro-cesses

involved

in

the

6

Li

production

and

destruction

are

given.

In

Section

4

a

short

description

of

the

LUNA

apparatus

is

reported

(more

details

in

[6]

).

In

Section

5

data

analysis

is

described

and

finally

in

Sections

6

and

7

the

2

H(

α

,

γ

)

6

Li

reaction

rate

includ-ing

the

new

LUNA

data

points

and

the

resulting

conclusions

are

given.

2.

Review

of

astronomical

data

on

6

Li

6

Li

is

mainly

produced

during

the

BBN

epoch

and,

in

more

recent

epochs,

by

cosmic

ray

spallation

[7]

.

For

this

reason,

its

primordial

abundance

is

inferred

from

observations

of

the

atmo-spheres

of

hot

metal-poor

stars

in

the

galactic

halo

(either

main

sequence

dwarfs

or

subgiants

near

the

turn-off point).

The

primor-dial

abundance

is

then

obtained

by

extrapolating

the

abundance

at

zero

metallicity.

The

strength

of

the

lithium

absorption

line

(

λ

=

670.7

nm)

provides

the

lithium

abundance.

As

the

absorption

line

of

6

Li

is

slightly

shifted

towards

a

higher

wavelength

compared

to

7

Li,

the

abundance

of

6

Li

isotope

can

be

derived

through

the

shape

analysis

of

the

lithium

absorption

line.

The

shape

of

the

absorption

line

is

also

affected

by

the

convective

motions

of

the

stellar

atmo-sphere

thus

analysis

of

the

lithium

absorption

line

depends

on

the

theoretical

model

for

the

stellar

atmosphere.

Recently,

Steffen

et

al.

[10]

and

Lind

et

al.

[11]

have

pointed

out

that,

by

using

three-dimensional

model

atmospheres

and

dropping

the

assumption

of

Local

Thermodynamic

Equilibrium

(3D

NLTE),

many

of

the

6

Li

detections

may

become

non-significant.

However,

the

situation

is

still

unclear.

As

an

example,

3D-NLTE

analysis

of

star

HD

84937

provides

contradictory

results

according

to

differ-ent

authors

[10,11]

.

In

other

cases,

such

as

star

G020-024,

we

have

a

good

agreement

between

one

dimensional

[3]

and

three

dimen-sional

models

[10]

,

respectively.

The

results

obtained

in

metal-poor

Table 1

Astronomical observations. Relative 6 Li/ 7 Li isotopic

abundance in some metal-poor stars obtained using dif- ferent stellar atmosphere models, in one (1D) or three (3D) dimensions, with or without Local Thermodynamic Equilibrium (LTE, NLTE), respectively.

Star name 6 Li/ 7 Li [%] method ref.

HD 84937 5 .0 ± 2.0 1D LTE [8] 5 .2 ± 1.9 1D LTE [9] 6 .6 ± 2.4 1D LTE [10] 5 .1 ± 1.5 1D NLTE [4] 5 .1 ± 2.3 3D NLTE [10] 1 .1 ± 1.0 3D NLTE [11] HD 160617 0 .1 ± 1.2 1D LTE [10] 0 .3 ± 1.2 1D LTE [10] 3 .6 ± 1.0 1D LTE/NLTE [3] −0.5 ± 1.2 3D NLTE [10] −1.3 ± 1.2 3D NLTE [10] −0.01 ± 0.01 3D NLTE [12] G064-012 5 .9 ± 2.1 1D NLTE [4] 2 .3 ± 2.9 1D LTE [10] 2 .0 ± 2.9 1D LTE [10] 1 .2 ± 2.8 3D NLTE [10] 0 .8 ± 2.7 3D NLTE [10] 0 .5 ± 2.3 3D NLTE [11] G020-024 11 .7 ± 2.5 1D LTE [10] 7 .0 ± 1.7 1D LTE [3] 9 .9 ± 2.4 3D LTE [10]

stars

(HD

84937,

HD

160617,

G064-012

and

G020-024)

where

dif-ferent

analyses

have

been

reported

in

the

literature,

are

shown

in

Table

1

.

In

conclusion,

although

several

authors

agree

on

the

fact

that

6

Li

has

been

detected

on

a

few

halo

stars

[2,10]

,

the

determination

of

primordial

6

Li

abundance

is

still

a

difficult

task,

because

obser-vations

are

affected

by

systematic

uncertainties

related

to

the

con-vective

motion

of

the

stellar

atmosphere.

As

a

consequence,

obser-vations

with

spectrometers

with

higher

sensitivity

and

resolution

would

be

necessary.

At

the

same

time,

it

would

be

important

to

improve

stellar

modeling

in

order

to

firmly

establish

the

primor-dial

abundance

of

6

Li

inferred

from

observations.

3.

The

nuclear

physics

of

primordial

6

Li

In

the

standard

BBN

framework,

the

primordial

6

Li

abundance

is

mainly

determined

by

two

nuclear

reactions

[13]

:

the

2

H(

α

,

γ

)

6

Li

reaction

that

produces

6

Li

and

the

6

Li(p,

α

)

3

He

that

destroys

it.

The

6

Li(p,

α

)

3

He

reaction

rate

is

fairly

well

known

in

the

BBN

energy

range

[14]

.

On

the

other

hand,

the

lack

of

direct

measurements

at

low

energy

of

the

2

H(

α

,

γ

)

6

Li

process

makes

its

reaction

rate

largely

uncertain.

The

2

H(

α

,

γ

)

6

Li

cross

section

at

energies

less

than

1

MeV

is

dominated

by

radiative

E2

capture

from

d

waves

in

the

scattering

state

into

the

ground

state

of

6

Li

through

a

3

+

resonance

at

E

R

=

711

keV.

At

energies

lower

than

300

keV

in

the

center-of-mass

sys-tem

system,

due

to

the

different

angular

momentum

barriers

for

p

and

d

waves,

the

E1

contribution

is

expected

to

be

comparable

to

the

E2

one

[15]

.

For

a

detailed

discussion

of

theoretical

predictions,

see

Refs.

[15,16]

.

The

2

H(

α

,

γ

)

6

Li

cross

section

has

never

been

directly

measured

in

the

BBN

region

of

interest

(30

E

300

keV),

before

the

LUNA

data

first

reported

in

[5]

.

It

is

worth

nothing,

only

one

attempt

to

directly

measure

the

2

H(

α

,

γ

)

6

Li

cross

section

at

low

energy

is

re-ported

in

literature,

providing

only

an

upper

limit

[17]

.

The

other

direct

measurements

have

been

performed

at

higher

energies,

i.e.

around

the

711

keV

resonance

[18]

,

and

in

the

MeV

region

[19]

.

In

order

to

unfold

the

2

H(

α

,

γ

)

6

Li

cross

section

at

low

energy,

two

dif-ferent

Coulomb

dissociation

experiments

have

been

performed

at

26

[20]

and

150

MeV/nucleon

[15]

projectile

energy,

respectively.

(4)

Fig. 1. Experimental setup. The 4 He + beam enters in the gas target through the

collimator (1), passes through the steel pipe (2) and reaches the calorimeter (3). The deuterium inlet (4) is below the calorimeter. The pressure is kept constant by means of a feedback system controlled by a pressure gauge (5). The

γ

-rays pro- duced by the 2 H(

α

,

γ

) 6 Li reaction are detected by a HPGe detector (6). The beam

induced background (see Section 4.2 for details), is monitored by a silicon detector (7). To reduce the environmental background, a lead castle (8) and an anti-radon box (9) surround the detector and gas target.

This

technique

is

mainly

sensitive

to

the

electric

quadrupole

(E2)

component

of

the

cross

section.

However,

the

competing

nuclear

breakup

process

dominates

at

both

beam

energies

[15,20]

.

There-fore,

these

data

may

be

interpreted

as

upper

limits

of

the

E2

con-tribution

to

the

cross

section.

In

conclusion,

the

6

Li

abundance

pre-dicted

by

the

BBN

theory

is

affected

by

large

uncertainties

because

of

the

lack

of

data

for

the

2

H

(

α

,

γ

)

6

Li

reaction.

In

this

context,

the

LUNA

measurement

discussed

in

this

paper

represents

a

significant

step

forward,

providing

the

first

direct

measurement

of

this

reac-tion

well

inside

the

BBN

energy

region.

4.

Experimental

setup

4.1.

Apparatus

description

The

measurement

was

performed

by

detecting

the

prompt

γ

rays

emitted

from

the

2

H(

α

,

γ

)

6

Li

reaction,

using

the

experimen-tal

setup

shown

in

Fig.

1

.

A

4

He

+

beam

generated

by

the

400

keV

LUNA

II

accelerator

[21]

passed

several

collimators

along

a

pow-erful

gas

pumping

system

to

enter

the

target

chamber,

in

which

a

deuterium

gas

pressure

of

0.3

mbar

was

kept

by

a

feedback

system.

Together

with

the

good

energy

stability

of

the

accelerator,

the

win-dowless

gas

target

(length

17.7

cm)

allows

for

an

accurate

calcula-tion

of

the

beam

energy

along

the

entire

beam

path.

The

stopping

power

inside

the

gas

target

is

quite

low

(0.18

keV/cm

at

400

keV

beam

energy).

The

beam

intensity

(typically

0.3

mA)

was

measured

by

a

beam

calorimeter

with

a

constant

temperature

gradient.

The

target

chamber

and

HPGe

detector

were

shielded

by

at

least

20

cm

of

lead

in

all

directions

and

the

setup

was

enclosed

in

an

anti-radon

box

flushed

with

nitrogen

to

ensure

a

very

low

and

stable

natural

background

(see

Fig.

2

).

The

γ

-rays

were

de-tected

by

a

large

(137%

relative

efficiency)

high-purity

germanium

(HPGe)

detector

placed

at

a

90

° angle

with

respect

to

the

ion

beam

direction

in

a

close

geometry,

to

reach

a

full-energy

peak

detection

efficiency

of

1.7%

at

1.6

MeV.

An

energy

resolution

of

2.8

keV

at

1.6

MeV

γ

-ray

energy

was

achieved.

A

more

detailed

description

of

the

setup

can

be

found

in

[6,22]

.

4.2.

Neutron

induced

background

2

H(

α

,

γ

)

6

Li

is

not

the

only

reaction

that

occurred

in

the

gas

tar-get.

Deuterons

can

be

elastically

scattered

by

the

incoming

alpha

Fig. 2. Black line: Natural background measured with a HPGe detector. Blue line: The 2 H(

α

,

γ

) 6 Li spectrum at E α= 360 keV and p

target = 0.3 mbar (natural back-

ground has been subtracted). Violet line: The 2 H(

α

,

γ

) 6 Li simulated spectrum at E

α

= 360 keV and p target = 0.3 mbar (scaled down). The arrows indicate the (n,n’

γ

)

lines. The triangular peaks are due to the interaction of neutrons with the ger- manium (see text). More details are reported in [6] . The red band indicates the

2 H(

α

,

γ

) 6 Li region of interest at E

α= 360 keV. The measured beam induced back-

ground at LUNA is about one order of magnitude less than the natural background expected on Earth surface experiments with the same setup [24] . (For interpreta- tion of the references to colour in this figure legend, the reader is referred to the web version of this article.)

particles,

receiving

a

kinetic

energy

that

enables

them

to

induce

the

2

H(

2

H,p)

3

H

and

2

H(

2

H,n)

3

He

reactions

with

the

surrounding

deuterons

in

the

gas.

Protons

from

the

2

H(

2

H,p)

3

H

reaction

were

monitored

with

a

silicon

detector

(see

Fig.

1

and

[6]

for

details).

Neutrons

from

the

2

H(

2

H,n)

3

He

reaction

can

interact

with

the

sur-rounding

materials

(iron,

copper,

lead

and

germanium)

producing

γ

-rays

through

(n,n’

γ

)

reactions.

In

order

to

reduce

the

deuteron

scattering

probability,

a

pipe

with

a

square

section

of

2

cm

was

in-stalled

along

the

beam

line.

This

pipe

limits

the

free

path

of

the

elastically

scattered

deuterons

inside

the

gas,

thus

reducing

the

number

of

neutrons

produced

by

the

2

H(

2

H,n)

3

He

reaction

to

a

value

of

about

10

neutrons

per

second.

The

rate

of

this

neutron

in-duced

background

BG

neutron

in

the

Region

of

Interest

(RoI)

for

the

2

H(

α

,

γ

)

6

Li

reaction

is

about

one

order

of

magnitude

higher

than

the

expected

signal

but

nevertheless

much

lower

than

the

usual

background

present

on

Earth’s

surface

experiments

[23]

.

Fig.

2

shows

the

in-beam

E

α

=

360

keV

γ

-ray

spectrum

as

well

as

a

comparative

background

spectrum.

It

contains

many

lines

and

several

triangular

shaped

peaks

(for

more

details

see

[6]

).

The

for-mers

are

due

to

the

natural

background

and

(n,n’

γ

)

beam

induced

reactions

on

the

materials

surrounding

the

reaction

chamber.

The

triangular

shaped

peaks

are

due

to

(n,n’

γ

)

beam

induced

reactions

of

energetic

neutrons

on

the

Germanium

detector

itself.

The

tri-angular

shape

arises

from

the

germanium

nucleus

recoil

energy

deposited

in

the

detector.

Similar

features

have

been

observed

at

all

beam

energies

investigated.

The

neutron

induced

background

spectrum

as

a

function

of

the

beam

energy

has

been

studied

with

a

dedicated

simulation,

in

which

our

experimental

setup

has

been

reproduced.

The

simulated

spectra

are

consistent

with

the

exper-imental

ones

[22]

.

The

main

difference

between

the

spectra

ob-tained

at

different

beam

energies

(experimental

and

simulated),

is

an

overall

energy

dependence

of

the

HPGe

response

at

E

γ

1500

keV,

in

such

a

way

the

bin-by-bin

ratio

produced

at

dif-ferent

beam

energies

is

not

constant

but

weakly

depends

on

the

beam

energy.

For

what

concerns

the

complex

spectral

structure,

the

most

prominent

difference

between

spectra

at

different

beam

(5)

60 D. Trezzi et al. / Astroparticle Physics 89 (2017) 57–65 0 500 1000 1500 2000 2500 3000 3500 4000 4500 1450 1500 1550 1600 a.u. Energy [keV] isotropic emission pure E1 emission pure E2 emission

Fig. 3. Simulated spectra of the 2 H(

α

,

γ

) 6 Li reaction (E α= 400 keV), assuming dif-

ferent angular distribution for the emitted photons. Note that the RoI does not de- pends on the adopted angular distributions.

energies

is

related

to

the

excitation

of

56

Fe

nuclei

to

the

2658

keV

state,

which

decays

in

a

cascade

via

emission

of

1811

keV

and

847

keV

γ

-rays.

The

strength

of

the

gamma

lines

associated

to

the

2658

keV

state

increases

with

the

beam

energy.

In

fact,

at

high

energy

(360

or

400

keV)

there

are

more

neutrons

available

to

pop-ulate

the

2658

keV

level

with

respect

to

low

beam

energies

(280

or

240

keV),

thus

explaining

the

increase

of

the

1811

keV

counting

rate

with

the

beam

energy.

This

effect

is

confirmed

in

a

dedicated

simulation

in

which

the

Rutherford

α

(

2

H,

2

H)

α

scattering

and

the

2

H(

2

H,n)

3

He

reaction

are

implemented.

As

expected,

it

has

been

found

that

the

doppler

effect

determines

a

broadening

of

neutron

energy

distribution

with

the

beam

energy,

in

agreement

with

the

observed

effect

on

the

56

Fe

gamma

lines

[6,22]

.

5.

Data

analysis

5.1.

General

approach

In

this

section

the

data

analysis

of

the

2

H(

α

,

γ

)

6

Li

reaction

at

beam

energies

240,

280,

360

and

400

keV

is

reported.

The

energy

of

photons

produced

in

the

2

H(

α

,

γ

)

6

Li

reaction

depends

on

the

beam

energy

and

on

the

doppler

effect.

It

can

be

expressed

by

the

following

relativistic

formula

(in

which

c

=

¯h

=

1

):

E

γ

=

m

2 He

+

m

2d

− m

2 Li

+

2

m

d

(

E

α

+

m

He

)

2

[

E

α

+

m

He

+

m

d

− p

He

cos

(

θ

lab

)

]

(1)

In

this

equation

E

γ

is

the

photon

energy,

m

He

,

m

d

and

m

Li

are

the

masses

of

nuclides

involved

in

the

reaction,

p

He

=



E

α

(

E

α

+

2

m

He

)

is

the

α

particle

momentum

and

θ

lab

is

the

angle

of

emitted

photon

with

respect

to

the

beam

axis.

This

formula

indicates

that

fully

detected

photons

populate

a

well

defined

region

of

interest

(RoI)

for

each

beam

energy.

The

energy

range

of

the

RoIs

only

depends

on

the

beam

energy

and

on

the

angular

acceptance

of

the

HPGe

detector

(10

°

θ

lab

170

°).

It

has

been

determined

with

the

GEANT

Monte

Carlo

sim-ulation,

in

which

the

experimental

conditions

are

reproduced

and

the

Eq.

(1)

is

implemented.

The

angular

distribution

of

photons

produced

by

the

2

H(

α

,

γ

)

6

Li

only

affects

the

distribution

of

fully

detected

photons

inside

the

RoI.

Fig.

3

shows

the

simulated

en-ergy

spectra

for

different

angular

distribution

of

emitted

pho-tons.

The

fully

detected

photons

produced

by

the

2

H(

α

,

γ

)

6

Li

re-action

have

the

following

RoIs:

[1542,1563]

keV,

[1554,1577]

keV,

[1580,1606]

keV

and

[1592,1620]

keV,

respectively

for

beam

ener-gies

of

240,

280,

360

and

400

keV.

We

developed

two

independent

analysis

methods.

Our

data

have

been

grouped

in

pairs

(i,j)

of

different

beam

energy

runs

char-acterized

by

non-overlapping

RoIs,

comparable

statistics

and

simi-lar

integrated

beam

current.

If

we

now

indicate

with

BG

neutron

i

and

BG

neutron

j

the

rate

of

neutron

induced

background

relative

to

i

and

j

energies,

we

have:

BG

neutron

i

=

β

i j

BG

neutronj

(2)

As

it

has

been

discussed

in

Section

4.2

,

the

shape

of

beam

in-duced

background

weakly

depends

on

the

beam

energy.

In

this

for-mula

this

dependence

is

included

in

the

β

ij

parameter

and

will

be

considered

in

the

following.

Both

analysis

methods

provide

cross

section

values

σ

(

E

),

pa-rameterized

by

the

astrophysical

S-factor

given

by:

S

(

E

)

=

σ

(

E

)

E

exp



72

.

44

/

E



(3)

where

E

(

keV

)

is

the

center-of-mass

system

energy.

The

two

analy-sis

methods

named

“Flat

regions” and

“Minimization” will

now

be

described

in

detail.

5.1.1.

Method

A,

based

on

“flat

regions”

After

subtraction

of

the

natural

background,

there

are

several

energy

regions

without

any

visible

peak.

This

purely

neutron-induced

background

is

mainly

due

to

the

Compton

continua

of

several

peaks

in

the

γ

-ray

spectrum.

Therefore,

it

is

possible

to

parameterize

the

ratio

of

background

rates

inside

such

“flat

re-gions” at

two

beam

energies

along

the

entire

γ

-ray

spectrum.

In

this

way,

the

ratio

of

background

levels

can

be

calculated

anywhere

in

the

γ

-ray

spectrum.

This

allows

to

properly

subtract

beam

in-duced

backgrounds

obtained

at

two

different

beam

energies

using

as

normalization

factor

the

background

ratio.

Given

that

the

two

RoIs

do

not

overlap,

The

two

signals

Y

Enetα=400 keV

(RoI

400

keV)

and

Y

Enetα=280 keV

(RoI

280

keV),

can

be

written

as

Y

400 net

(

400

)

=

Y

400

(

400

)

β

400,280

Y

400

(

280

)

(4)

Y

280 net

(

280

)

=

Y

280

(

280

)

1

β

400,280

Y

280

(

400

)

(5)

Similar

equations

can

be

written

for

the

other

two

beam

en-ergies

(E

α

=

240

and

360

keV).

The

systematic

uncertainty

com-ing

from

the

determination

of

the

normalization

factor

(here

called

β

ROI

400,280

and

β

360ROI,240

) from neighboring “flat regions” is difficult to

estimate.

Therefore,

instead

of

deriving

the

systematic

error

bar

of

β

ROI

400,280

,

an

approach

is

chosen

that

cancels

its

effects

on

the

sig-nal

yield,

and

thus

on

the

final

astrophysical

S-factor.

It

is

clear

from

Eqs.

(4)

and

(5)

that

β

ROI

400,280

uncertainties

affect

the

two

yields

in

opposite

ways:

If

β

ROI

400,280

is

overestimated,

Y

net400

(

400

)

de-creases

but

Y

280

net

(

280

)

increases.

However,

we

found

that

the

yield

sum

Y

400

net

(

400

)

+

Y

net280

(

280

)

is

independent

from

β

400ROI,280

,

so

that

the

issue

of

the

error

estimation

of

this

parameter

becomes

moot.

To

extract

the

astrophysical

S-factor

at

a

given

energy

from

the

yield

sum

Y

400

net

(

400

)

+

Y

net280

(

280

)

,

it

is

necessary

to

make

an

as-sumption

on

the

energy

dependence

of

the

astrophysical

S-factor

inside

the

LUNA

energy

range

[25]

.

For

the

data

analysis,

various

assumptions

ranging

from

S

(400

keV)/

S

(280

keV)

=

1.1–1.5

have

been

tested,

where

1.1

corresponds

to

pure

E1

and

1.5

to

pure

E2.

The

S-factor

result

does

not

change

outside

its

statistical

error

bar

in

any

of

the

cases

tested.

The

Method

A

analysis

results

are

shown

in

Table

2

,

where

the

theoretical

S-factor

trend

indicated

in

[15]

is

assumed,

with

a

ratio

S

(400

keV)/

S

(280

keV)

=

1.3.

In

summary,

in

method

A

the

effects

of

the

background

subtrac-tion

cancel

out,

so

the

result

does

not

depend

on

the

uncertainty

of

the

background

subtraction.

However,

this

method

requires

a

the-oretical

assumption

for

the

shape

of

the

S-factor

curve,

and

several

(6)

Table 2

Astrophysical S-factor and nuclear cross section. In the table are indicated: the effective center-of-mass energy E [keV], the beam energy E α[keV], the deuterium gas target pressure P [mbar], the measurement time t [h], the integrated beam current Q [C] =

I beam t, the net reaction yield Y [counts/C] and its statistical error (the counting error for the “method A” includes the uncertainty

due to the theoretical assumption, see text), the total cross section

σ

[pb] with counting and systematic errors, the astrophysical S-factor S [keV

μ

b] with counting and systematic errors. The results of analysis “Method B” are adopted in this work. The data points at 80 and 93 keV are less than two standard deviations above zero and should be considered as upper limits.

E [keV] E α[keV] P [mbar] Q [C] t [h] Yield [counts/C] cross section [pb] S-factor [keV

μ

b] Method 80 240 0 .306 211 .5 217 .7 0 .23 ± 0.64 9 ± 25 ± 1 2 .4 ± 6.7 ± 0.3 A 93 280 0 .308 538 .9 490 .9 0 .46 ± 0.46 18 ± 18 ± 2 3 .0 ± 3.0 ± 0.4 A 120 360 0 .306 252 .7 205 .2 0 .94 ∓ 0.57 37 ∓ 23 ± 5 3 .3 ∓ 2.0 ± 0.4 A 133 400 0 .306 514 .3 437 .7 1 .46 ∓ 0.46 58 ∓ 18 ± 8 4 .1 ∓ 1.3 ± 0.5 A 80 240 0 .306 211 .5 217 .7 0 .11 ± 0.51 4 ± 20 ± 0 .4 1 .1 ± 5.2 ± 0.1 B 93 280 0 .308 538 .9 490 .9 0 .42 ± 0.25 16 ± 10 ± 2 2 .7 ± 1.6 ± 0.3 B 120 360 0 .306 252 .7 205 .2 1 .05 ± 0.47 40 ± 18 ± 6 3 .6 ± 1.6 ± 0.5 B 133 400 0 .306 514 .3 437 .7 1 .50 ± 0.32 59 ± 13 ± 7 4 .1 ± 0.9 ± 0.5 B

assumptions

have

been

tested.

Even

still,

a

radically

different

S-factor

curve

cannot,

in

principle,

be

excluded.

An

alternative

anal-ysis

method

to

derive

the

cross

section

without

any

theoretical

as-sumption

on

the

shape

of

the

S-factor

curve

is

described

in

the

next

section.

5.1.2.

Method

B,

based

on

minimization

procedures

This

method

is

based

on

Eq.

(2

).

The

γ

-ray

spectrum

rate

R

i

is

composed

by

the

neutron-induced

background

BG

neutron

i

,

the

natu-ral

background

BG

room

and

the

2

H(

α

,

γ

)

6

Li

counting

rate.

This

latter

can

be

expressed

as

k

i

N

i

,

in

which

k

i

is

a

scaling

pa-rameter

and

N

i

is

the

energy

distribution

due

to

the

photons

pro-duced

by

the

2

H(

α

,

γ

)

6

Li

reaction.

The

N

i

distribution

has

been

ob-tained

with

the

dedicated

GEANT

simulation

in

which

all

the

de-tails

of

the

experimental

conditions

are

reproduced

and

2

H(

α

,

γ

)

6

Li

reactions

are

generated

along

the

beam

axis

(see

Fig.

3

).

The

BG

neutron

rate

can

be

written

as

follows:

BG

neutron

i

=

R

i

− BG

room

− k

i

N

i

(6)

Substituting

Eq.

(6)

in

Eq.

(2)

we

obtain:

R

i

− BG

room

− k

i

N

i

β

i j



R

j

− BG

room

− k

j

N

j



=

0

(7)

With

the

“Method

B” the

free

parameters

β

ij

,

k

i

and

k

j

in

the

Eq.

(7)

are

obtained

with

a

MINUIT

χ

2

(least

squares)

minimiza-tion

procedure

in

the

energy

window

[1500,1625]

keV,

which

in-cludes

all

the

considered

RoIs.

As

stated

above,

the

shape

of

the

distribution

of

fully

detected

photons

inside

a

RoI

strongly

depends

on

the

adopted

angular

distribution

of

emitted

photons

(see

Fig.

3

).

In

this

concern,

all

the

counting

levels

inside

of

the

RoI’s

(natural

background,

beam

induced

background

and

signals)

have

been

averaged

before

the

minimization

procedure,

to

make

analysis

sensitive

only

to

the

to-tal

cross

section

and

not

dependent

on

angular

distribution

as-sumptions.

The

HPGe

efficiency

for

different

angular

distributions

has

been

tested

with

the

simulation.

As

a

result,

the

efficiency

un-certainty

due

to

the

unknown

angular

distribution

has

been

con-servatively

quoted

at

the

9%

level.

As

discussed

in

Section

4.2

,

the

β

ij

parameter

weakly

depends

on

the

energy.

This

dependence

has

been

modeled

with

a

second

order

polynomial

in

our

minimization

region.

This

correction

de-termines

a

negligible

variation

of

the

cross

sections,

about

a

fac-tor

four

lower

than

the

statistical

errors,

for

all

the

four

consid-ered

data

points.

As

discussed

in

Section

4.2

,

the

strength

of

the

1811

keV

line

slightly

depends

on

the

beam

energy.

Although

the

line

is

outside

the

minimization

interval,

its

Compton

edge

is

at

about

1580

keV.

To

evaluate

its

effect

in

the

analysis

results,

this

line

and

its

Compton

edge

have

been

subtracted

from

the

exper-imental

spectra.

To

do

so,

the

strength

of

the

1811

keV

peak

of

each

dataset

has

been

used

to

normalize

and

subtract

the

simu-lated

spectrum.

This

correction

was

found

to

be

negligible

as

well,

about

a

factor

20

lower

than

the

statistical

error,

for

all

the

four

energies

considered.

The

results

of

Method

B

described

above

are

shown

in

Table

2

.

The

analysis

has

been

repeated

using

wider

en-ergy

windows

for

the

minimization

procedure,

up

to

500

<

E

<

2500

keV,

obtaining

results

fully

consistent

with

those

presented

in

Table

2

.

As

a

consistency

check,

the

minimization

procedure

has

been

repeated

in

energy

windows

not

containing

the

RoI’s.

As

ex-pected,

the

counting

difference

between

paired

spectra

is

always

consistent

with

zero,

within

statistical

fluctuations

[26]

.

6.

Results

and

astrophysical

aspects

6.1.

Astrophysical

S-factor

and

nuclear

cross

section

The

relevant

measurement

parameters

and

the

results

of

the

two

analysis

are

reported

in

Table

2

.

The

yield

Y

i

inside

a

RoI

is

defined

by

the

following

formula:

Y

i

=

k

i

n

RoIi

t

i

/Q

i

(8)

where

n

RoI

i

are

the

net

counts/sec

in

the

RoI,

t

i

is

the

acquisition

time.

In

the

“Method

A” the

error

of

Y

i

reported

in

Table

2

is

due

to

the

statistical

fluctuations

inside

the

two

paired

RoI’s

and

to

the

uncertainties

of

the

theoretical

assumption

adopted

to

estab-lish

the

energy

trend

of

the

cross

section.

The

yield

uncertainty

obtained

with

“Method

B” has

been

obtained

in

the

χ

2

(least

squares)

minimization

procedure

of

MINUIT,

in

which

the

corre-lation

between

β

,

k

i

,

and

k

j

is

taken

into

account

by

means

of

co-variance

matrix.

It

is

essentially

due

to

the

counting

fluctuations

and

therefore

it

determines

the

statistical

significance

of

the

ob-tained

results.

Note

that

the

significance

of

Y

400

net

(

400

)

is

more

than

4

σ

,

while

Y

240

net

(

240

)

is

compatible

with

zero.

As

a

matter

of

fact,

in

the

[1592,1620]

keV

RoI

(

E

cm

=

133

keV)

the

measured

excess

is

about

750

counts

and

the

background

level

is

10

times

higher

for

E

α

=

400,280

keV.

On

the

contrary,

the

excess

in

the

[1542,1563]

keV

RoI

(Ecm

=

80

keV)

is

of

only

23

events,

a

negligible

value

with

respect

to

the

statistical

fluctuation

of

the

corresponding

back-ground

(see

Table

2

).

For

both

the

analysis

the

systematic

error

depends

on

the

un-certainties

on

the

working

conditions

and

represents

a

possible

common

bias

to

all

the

measurements,

without

affecting

the

sta-tistical

significance

of

the

obtained

results.

Table

3

lists

the

main

sources

of

systematic

uncertainties.

The

largest

contribution

is

due

to

the

unknown

angular

distribution

of

the

emitted

γ

-rays,

affect-ing

the

percentage

of

detected

photons.

The

γ

-ray

detection

effi-ciency

was

measured

with

calibrated

137

Cs

,

60

Co

,

and

88

Y

sources

as

a

function

of

their

position

along

the

beam

axis.

Due

to

the

close

geometry

used,

the

data

have

been

corrected

for

the

true-coincidence

summing

out

effect

[6,22]

.

The

reproducibility

of

the

calorimeter

calibration

was

at

the

3%

level.

This

value

was

adopted

as

the

uncertainty

on

the

beam

Riferimenti

Documenti correlati

Indian and Chinese managers choose generally different leadership styles but when it comes to acquire companies in mature markets they use the same strategies if the purpose

delle disposizioni del titolo I-bis del t.u.f., ma legittima la stessa Commissione a compiere tutti gli atti necessari all’accertamento delle violazioni, utilizzando

the slow workings of the Roman Congregations of Index and Inquisition, Pius X envisioned new ways of control also for the church around the world in his fight against

Insomma, in età romano-imperiale, tutto il settore nordorientale della penisola, ordinato da Augusto come X regio e riorganizzato da Diocleziano come VIII

Quanti barioni ci possono

In fact Uccellacci e uccellini is the only film of its kind, but a detailed analysis of each element of Pasolini’s neologism will show how rooted into Pasolini’s artistic practice

A preponderant divergence between vinifera and sylvestris has been observed by several authors [38,44,45] but some studies have also identified a genetic signal that suggests

15 In particular, data can be processed without the consent of the data subject: (a) for the performance of a contract to which the data subject is party or in order to take steps